Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 29
Filter
Add more filters










Publication year range
1.
J Sep Sci ; 46(6): e2200827, 2023 Mar.
Article in English | MEDLINE | ID: mdl-36670088

ABSTRACT

Abscisic acid (2-cis,4-trans-abscisic acid) is a plant hormone that has an asymmetric carbon atom. We tried to separate the enantiomers of native abscisic acid by HPLC using a phenyl column and a chiral mobile phase containing γ-cyclodextrin. The optimum mobile phase conditions were found to be 0.8% (w/v) γ-cyclodextrin, 4% (v/v) acetonitrile, and 20 mM phosphate buffer (pH 6.0). It was found that (R)-abscisic acid was earlier detected than (S)-abscisic acid. Since γ-cyclodextrin is hardly retained on a phenyl column, it was suggested that (R)-abscisic acid formed a more stable complex with γ-cyclodextrin than the (S)-abscisic acid. Abscisic acid in an acacia honey sample was successfully enantioseparated with the proposed method and only (S)-abscisic acid was detected. A biologically inactive 2-trans,4-trans-abscisic acid, which was prepared by irradiation of abscisic acid with a light-emitting diode lamp at 365 nm, was partially enantioseparated by the proposed method. Since the irradiation of (S)-abscisic acid-induced cis-to-trans isomerization to produce one 2-trans,4-trans-abscisic acid enantiomer, it is reasonable that racemization did not proceed during the cis-to-trans isomerization. (S)-Abscisic acid and probably (S)-2-trans,4-trans-abscisic acid were detected in a honey sample, where the peak area of (S)-abscisic acid was 7 times larger than that of (S)-2-trans,4-trans-abscisic acid.


Subject(s)
beta-Cyclodextrins , gamma-Cyclodextrins , beta-Cyclodextrins/chemistry , Chromatography, High Pressure Liquid/methods , Abscisic Acid , Stereoisomerism , Indicators and Reagents
2.
J Chromatogr A ; 1673: 463029, 2022 Jun 21.
Article in English | MEDLINE | ID: mdl-35468371

ABSTRACT

Catechin and epicatechin were enantioseparated by high-performance liquid chromatography (HPLC) with a phenyl column and aqueous mobile phases containing 0.05% (w/v) and 0.6% (w/v) of ß-cyclodextrin for catechin and epicatechin, respectively. ß-Cyclodextrin was found to be scarcely retained on a phenyl column. Consequently, it was suggested that catechin, which was eluted earlier than epicatechin, formed more stable inclusion complex with ß-cyclodextrin than epicatechin and earlier eluted enantiomers, (-)-catechin and (+)-epicatechin, formed more stable diastereomer complexes with ß-cyclodextrin than the respective enantiomers. This was confirmed by ß-cyclodextrin-modified micellar electrokinetic chromatography and Benesi-Hildebrand plots by fluorescence spectrophotometry. Effect of sugars (D-sucrose, D-glucose, and D-fructose) on the epimerization of (+)-catechin and (+)-epicatechin by heating was investigated by HPLC with a ß-cyclodextrin stepwise elution mode, in which two kinds of aqueous eluents containing different concentrations of ß-cyclodextrin were used by turns. The epimerization of the two enantiomers was suppressed only when D-fructose was added. Separation of ten kinds of catechins including catechin and epicatechin enantiomers was investigated by a ß-cyclodextrin linear gradient HPLC elution mode without using organic solvents, where two kinds of aqueous eluents containing different concentrations of ß-cyclodextrin were used with changing their ratio gradually. These catechins in a green tea infusion could be separated successfully by this method.


Subject(s)
Catechin , Cyclodextrins , beta-Cyclodextrins , Catechin/analysis , Chromatography, High Pressure Liquid/methods , Cyclodextrins/chemistry , Fructose , Tea/chemistry , beta-Cyclodextrins/chemistry
3.
J Sep Sci ; 44(15): 2932-2940, 2021 Aug.
Article in English | MEDLINE | ID: mdl-34077621

ABSTRACT

Cyclodextrins and their derivatives have been used for chiral high-performance liquid chromatography selectors, while they are costly to use as mobile phase additives in high-performance liquid chromatography. Here, we report application of phenyl column coated permanently with methylated ß-cyclodextrin for chiral high-performance liquid chromatography. A 0.1% (v/v) phosphoric acid solution containing 1 M NaCl and 0.5% (w/v) methylated ß-cyclodextrin was subjected to a phenyl column at a flow rate of 0.5 mL/min at 30°C for 2 h. Using the precoating phenyl column, all the enantiomers of the four phenethylamines (norepinephrine, epinephrine, octopamine, and synephrine) were successfully separated simultaneously by high-performance liquid chromatography with a mobile phase without methylated ß-cyclodextrin at a flow rate of 0.2 mL/min at 30°C. The enantioseparation ability was retained for successive analyses during 1 week. It is suggested that inclusion complex of methylated ß-cyclodextrin with a phenyl group on the surface of the stationary phase could be formed and that the inclusion complex could form the ternary complex with the injected analytes. The longer retention time of (S)-enantiomers of analytes than corresponding (R)-enantiomers for high-performance liquid chromatography could be explained from the higher stability of the methylated ß-cyclodextrin complexes with (S)-enantiomers, which were confirmed by capillary electrophoresis and 1 H NMR spectroscopy experiments.


Subject(s)
Chromatography, High Pressure Liquid/methods , Phenethylamines/isolation & purification , beta-Cyclodextrins/chemistry , Electrophoresis, Capillary/methods , Methylation , Phenethylamines/chemistry , Proton Magnetic Resonance Spectroscopy/methods , Stereoisomerism
4.
Magn Reson Chem ; 58(10): 941-948, 2020 10.
Article in English | MEDLINE | ID: mdl-32510643

ABSTRACT

Because choice of chiral nuclear magnetic resonance (NMR) shift reagents and concentration conditions have been made empirically by trials and errors for chiral NMR analyses, the prediction of NMR signal separation behavior is an urgent issue. In this study, the separation of enantiomeric and enantiotopic 1 H and 13 C NMR signals for α-amino acids and tartaric acid was performed by using the praseodymium(III) complex with (S,S)-ethylenediamine-N,N'-disuccinate ((S,S)-EDDS). All the present D-amino acids exhibited larger downfield shift of their α-protons and α-carbons compared with those for the corresponding L-amino acids in common. This regularity is applicable to absolute configurational assignment and determination of optical purity of amino acids. The chemical shifts of ß-protons of d- and l-alanine fully bound with the Pr(III) ((S,S)-EDDS) complex (δb s) and the adduct formation constants of both enantiomers (Ks) were obtained by dependences of the observed downfield shifts of the ß-protons on the total concentrations of the respective enantiomers in the presence of a constant concentration of the Pr(III) complex. The difference in the K values was found to be predominant determining factor for the enantiomeric signal separation. The chemical shifts of both enantiomers (δs) and the enantiomeric signal separations (Δδs) under given conditions could be calculated from the δb and K values. Furthermore, prediction of the signal separation behavior was enabled by using the calculated δ values and the signal broadening obtained by dependences of the half-height widths of the observed signals on the bound/free substrate concentration ratios for the respective enantiomers.

5.
J Oleo Sci ; 69(7): 677-684, 2020 Jul 02.
Article in English | MEDLINE | ID: mdl-32522947

ABSTRACT

A simple screening method for discrimination between commercial extra virgin olive oils and their blends with other vegetable oils was developed. Squalene, which was contained relatively high amounts in virgin olive oil, was determined by HPLC after a simple pretreatment that was carried out by dilution of oil samples with 2-propanol. Tyrosol, which was contained at relatively high concentration in virgin olive oil among phenolic compounds, was determined by HPLC after a simple liquid-liquid extraction. When using squalene and tyrosol contents as axes, extra virgin olive oils could be discriminated from pure olive oils, blended oils (extra virgin olive oils with sunflower oil or grapeseed oil) and other vegetable oils. These results suggest that determining squalene and tyrosol in seed oil samples could be useful in distinguishing between extra virgin olive oil and blended oils as a screening method.


Subject(s)
Food Analysis/methods , Food Contamination/analysis , Food Contamination/prevention & control , Food Quality , Olive Oil/analysis , Olive Oil/chemistry , Phenylethyl Alcohol/analogs & derivatives , Squalene/analysis , Chromatography, High Pressure Liquid , Liquid-Liquid Extraction/methods , Phenylethyl Alcohol/analysis , Plant Oils/analysis
6.
Chirality ; 32(7): 1020-1029, 2020 07.
Article in English | MEDLINE | ID: mdl-32346918

ABSTRACT

Direct enantioseparation of mandelic acid by high-performance liquid chromatography (HPLC) with a reversed phase column and a mobile phase containing a small amount of hydroxylpropyl-ß-cyclodextrin (HP-ß-CD) was studied as an efficient method for saving consumption of the CD additive. As a result, it was proposed that racemic mandelic acid can be analyzed with a phenyl column by using a mobile phase composed of 10 mM ammonium acetate buffer (pH 4.2) and 0.02% (w/v) HP-ß-CD at a flow rate of 1.0 mL/min at 40°C after the passage of 10 mM ammonium acetate buffer (pH 4.2) containing 0.1% (w/v) HP-ß-CD as a precoating mobile phase for 60 min. It is suggested that HP-ß-CD is bound with a phenyl group on the surface of the stationary phase to allow a phenyl column to act as a transient chiral column, and injected mandelic acid can form the ternary complex with the adsorbed HP-ß-CD. The longer retention time of D-mandelic acid than the L-isomer for HPLC can be explained from the higher stability of the HP-ß-CD complex with D-mandelic acid, which was confirmed by CE experiment with HP-ß-CD as a selector. The efficiency of a phenyl column compared with other stationary phases was also discussed.

7.
RSC Adv ; 9(43): 25177-25183, 2019 Aug 08.
Article in English | MEDLINE | ID: mdl-35528666

ABSTRACT

The reactions of a dichloro(p-cymene)ruthenium(ii) dimer, [RuCl2(p-cymene)]2, with citric acid and sulfur-containing amino acids gave only [Ru(L)(p-cymene)]-type complexes (L = citrate (Cit), l-penicillaminate (l-Pen), S-methyl-l-cysteinate (S-Me-l-Cys) and l-methioninate (l-Met)) in aqueous solutions at various pHs and molar ratios of the reactants, where Cit and the amino acids act as a tridentate ligand. These sulfur-containing amino acid complexes with bound nitrogen, oxygen and sulfur atoms and η6-p-cymene take S absolute configuration around Ru(ii) selectively, having the α-proton oriented in the opposite direction from the Ru(ii) center. The concentration dependences of the observed pseudo-first-order rate constants were provided for the substitution reactions of the citrato complex, [Ru(Cit)(p-cymene)], with a large excess of the sulfur-containing amino acids at various temperatures at pH 7.3, where solvolysis path was observed for S-Me-l-Cys and l-Met as an intercept but not for l-Pen. The activation parameters for the substitution reactions by the direct attack of the amino acids were changed significantly, indicating that the reaction mechanism varies sensitively with the amino acids from an associative mechanism to an interchange one. The pH dependences of the rate constants of the substitution reactions suggest that the carboxylate group is an attacking group for S-Me-l-Cys and l-Met under neutral conditions and the thiol group of l-Pen acts as an entering group constantly at any pH showing a considerably smaller activation energy compared with S-Me-l-Cys and l-Met. Differences in stabilities of the amino acid complexes were obtained from the equilibrium constants for the substitution reactions between the amino acids. These results indicate that the activation energies for the substitution reactions of the citrato complex with the amino acids are moderately correlated with the stabilities of the formed amino acid complexes.

8.
Anal Sci ; 35(4): 407-412, 2019 Apr 10.
Article in English | MEDLINE | ID: mdl-30555107

ABSTRACT

Racemic synephrine, which was transformed into diastereomers by derivatization with 2,3,4,6-tetra-O-acetyl-ß-D-glucopyranosil isothiocyanate, was resolved by a reversed phase HPLC with UV detection at 254 nm. The total contents of synephrine enantiomers in citrus fruit samples were exocarp > mesocarp > endocarp > sarcocarp, suggesting that synephrine content of outer side of citrus fruits was higher than that of the inner side. (R)-Synephrine was detected in exocarp of eleven fresh citrus fruits, except for lemon, lime, and grapefruit samples. (S)-Synephrine was determined in the exocarp of four citrus fruits (mikan, orange, bitter orange, and ponkan samples) and the ratio of (S)-synephrine to total synephrine was 0.5 - 0.9%. The racemization of (R)-synephrine in aqueous solution during heating at 100°C was also examined. An increase in the heating time brought about an increase in the (S)-synephrine content in a linear fashion. The racemization was found to be significantly reduced by the addition of D-fructose, D-maltose, D-glucose, D-mannose or D-galactose, but not D-sucrose or D-mannitol. It is suggested that the reducibility of sugars may result in the inhibition of racemization.


Subject(s)
Citrus/chemistry , Fruit/chemistry , Plant Extracts/isolation & purification , Synephrine/isolation & purification , Chromatography, High Pressure Liquid , Molecular Structure , Plant Extracts/chemistry , Stereoisomerism , Synephrine/chemistry
9.
Biomed Chromatogr ; 32(10): e4289, 2018 Oct.
Article in English | MEDLINE | ID: mdl-29768665

ABSTRACT

Xanthohumol, isoxanthohumol, and 8-prenylnaringenin in beer, hop and hop pellet samples were analyzed by HPLC using an InertSustain phenyl column and the mobile phase containing 40% methanol and 12% 2-propanol. Fractions of isoxanthohumol and 8-prenylnaringenin obtained by the above HPLC were separately collected. Isoxanthohumol and 8-prenylnaringenin were enantioseparated by HPLC using a Chiralcel OD-H column with a mobile phase composed of hexane-ethanol (90:10, v/v) and a Chiralpak AD-RH column with a mobile phase composed of methanol-2-propanol-water (40:20:40, v/v/v), respectively. Isoxanthohumol and 8-prenylnaringenin from beer, hop and hop pellet samples were found to be present in a racemic mixture. This can be explained by the fact that the two analytes were produced by a nonenzymatic process. The effects of boiling conditions on the conversion of xanthohumol into isoxanthohumol were also studied. A higher concentration of ethanol in heating solvent resulted in a decrease in the conversion ratio and the conversion was stopped by addition of ethanol at >50% (v/v). The isomerization was significantly affected pH (2-10) and the boiling medium at pH 5 was minimum for the conversion. Therefore, it was suggested that xanthohumol was relatively difficult to convert to isoxanthohumol in wort (pH 5-5.5) during boiling.


Subject(s)
Beer/analysis , Chromatography, High Pressure Liquid/methods , Flavanones/isolation & purification , Xanthones/isolation & purification , Flavanones/analysis , Flavanones/chemistry , Humulus/chemistry , Limit of Detection , Linear Models , Reproducibility of Results , Stereoisomerism , Xanthones/analysis , Xanthones/chemistry
10.
J Sep Sci ; 40(21): 4168-4175, 2017 Nov.
Article in English | MEDLINE | ID: mdl-28851084

ABSTRACT

We developed a reversed-phase high-performance liquid chromatography method with ultraviolet detection using on-line complexation with Cu(II) ion for analysis of five alcohols including diols and triol (methanol, ethanol, 1,2-propanediol, 1,3-propanediol, and glycerol). The Cu(II) ion concentration in the mobile phase had a great influence on the peak areas of these alcohols, but not on their retention times. Column temperature (25-40°C) and pH of the mobile phase did not affect the separation of analytes. The optimum separation conditions were determined as 5 mM CuSO4 , 3 mM H2 SO4 , and 3 mM NaOH at 30°C. The ratio of the peak areas for three alcohols (methanol, 1,2-propanediol, and glycerol) was in good agreement with that calculated from the obtained stability constants, molar absorption coefficients for the 1:1 Cu(II) complexes with the three alcohols, and the injected molar quantities. This fact strongly suggests that the observed high-performance liquid chromatography signals resulted from formation of the 1:1 Cu(II)-alcohol complexes. Using the proposed method, these five alcohols in spirit, liquid for electronic cigarette, mouthwash, and nail enamel remover samples were successfully analyzed with only a simple pretreatment.


Subject(s)
Alcohols/analysis , Copper , Cosmetics/analysis , Electronic Nicotine Delivery Systems , Mouthwashes/analysis , Chromatography, High Pressure Liquid , Chromatography, Reverse-Phase
11.
Chirality ; 29(6): 273-281, 2017 Jun.
Article in English | MEDLINE | ID: mdl-28403536

ABSTRACT

Enantiomeric 1 H and 13 C NMR signal separation behaviors of various α-amino acids and DL-tartarate were investigated by using the samarium(III) and neodymium(III) complexes with (S,S)-ethylenediamine-N,N'-disuccinate as chiral shift reagents. A relatively smaller concentration ratio of the lanthanide(III) complex to substrates was suitable for the neodymium(III) complex compared with the samarium(III) one, striking a balance between relatively greater signal separation and broadening. To clarify the difference in the signal separation behavior, the chemical shifts of ß-protons for fully bound D- and L-alanine (δb (D) and δb (L)) and their adduct formation constants (Ks) were obtained for both metal complexes. Preference for D-alanine was similarly observed for both complexes, while it was revealed that the difference between the δb (D) and δb (L) values is the significant factor to determine the enantiomeric signal separation. The neodymium(III) and samarium(III) complexes can be used complementarily for higher and smaller concentration ranges of substrates, respectively, because the neodymium(III) complex gives the larger difference between the δb (D) and δb (L) values with greater signal broadening compared to the samarium(III) complex.


Subject(s)
Ethylenediamines/chemistry , Neodymium/chemistry , Organometallic Compounds/chemistry , Samarium/chemistry , Ligands , Magnetic Resonance Spectroscopy , Stereoisomerism
12.
J Agric Food Chem ; 65(1): 244-250, 2017 Jan 11.
Article in English | MEDLINE | ID: mdl-27989117

ABSTRACT

Racemic sulforaphane, which was derivatized with (S)-leucine (l-leucine), was resolved by reversed phase HPLC with UV detection. The optimum mobile phase conditions were found to be 10 mM citric acid (pH 2.8) containing 22% methanol at 35 °C using detection at 254 nm. Sulforaphane enantiomers in florets and stems of five brands of broccoli and leaves and stems of three brands of broccoli sprouts were analyzed by the proposed HPLC method. Both sulforaphane enantiomers were detected in all of the samples. The S/R ratios of sulforaphane in broccoli samples were 1.5-2.6/97.4-98.5% for florets and 5.0-12.1/87.9-95.0% for stems. The S/R ratios in broccoli sprout samples were higher than those in broccoli samples and were found to be 8.3-19.7/80.3-91.7% for leaves and 37.0-41.8/58.2-63.0% for stems. (S)-Sulforaphane detected in the broccoli and its sprout samples was positively identified by separately using an HPLC with a chiral column (Chiralpak AD-RH) and mass spectrometry.


Subject(s)
Brassica/chemistry , Chromatography, High Pressure Liquid/methods , Isothiocyanates/chemistry , Leucine/chemistry , Brassica/growth & development , Isothiocyanates/isolation & purification , Plant Leaves/chemistry , Plant Leaves/growth & development , Plant Stems/chemistry , Plant Stems/growth & development , Seedlings/chemistry , Seedlings/growth & development , Stereoisomerism , Sulfoxides
13.
Chirality ; 28(1): 85-91, 2016 Jan.
Article in English | MEDLINE | ID: mdl-26512870

ABSTRACT

The optically active mixed-ligand fac(S)-tris(thiolato)rhodium(III) complexes, ΔL -fac(S)-[Rh(aet)2 (L-cys-N,S)](-) (aet = 2-aminoethanethiolate, L-cys = L-cysteinate) () and ΔLL -fac(S)-[Rh(aet)(L-cys-N,S)2 ](2-) were newly prepared by the equatorial preference of the carboxyl group in the coordinated L-cys ligand. The amide formation reaction of with 1,10-diaminodecane and polyallylamine gave the diamine-bridged dinuclear Rh(III) complex and the single-chain polymer-supported Rh(III) complex with retention of the ΔL configuration of , respectively. These Rh(III) complexes reacted with Co(III) or Co(II) to give the linear-type trinuclear structure with the S-bridged Co(III) center and the two Δ-Rh(III) terminal moieties. The polymer-supported Rh(III) complex was applied not only to the CD spectropolarimetric detection and determination of a trace of precious metal ions such as Au(III), Pt(II), and Pd(II) but also to concentration and extraction of these metal ions into the solid polymer phase. Chirality 28:85-91, 2016. © 2015 Wiley Periodicals, Inc.

14.
Anal Sci ; 31(8): 831-5, 2015.
Article in English | MEDLINE | ID: mdl-26256608

ABSTRACT

In order to analyze trigonelline, caffeine, chlorogenic acid, and their related compounds simultaneously, an HPLC method using an InertSustain C18 column and a mobile phase containing octanesulfonate as an ion-pairing reagent under an acidic condition was developed. The optimum mobile phase conditions were determined to be 0.1% phosphoric acid, 4 mM octanesulfonate, and 15% methanol at 35°C. Using the proposed method, trigonelline, nicotinic acid, caffeine, theophylline, chlorogenic acid, and caffeic acid in ten instant coffee samples were analyzed. These analytes except for theophylline were detected in all samples. An increase in the caffeine content in instant coffee samples tended to decrease in both trigonelline and chlorogenic acid contents, and the trigonelline content was found to be correlated well with the chlorogenic acid content (R(2) = 0.887).


Subject(s)
Alkaloids/analysis , Alkanesulfonic Acids/chemistry , Caffeine/analysis , Chlorogenic Acid/analysis , Chromatography, High Pressure Liquid/methods , Coffee/chemistry , Alkaloids/chemistry , Alkaloids/isolation & purification , Caffeine/chemistry , Caffeine/isolation & purification , Chlorogenic Acid/chemistry , Chlorogenic Acid/isolation & purification , Time Factors
15.
Chirality ; 27(7): 417-21, 2015 Jul.
Article in English | MEDLINE | ID: mdl-25994510

ABSTRACT

Three aldohexoses, glucose, galactose, and mannose, and three aldopentoses, arabinose, xylose, and ribose, were derivatized with L-tryptophanamide (L-TrpNH2 ) under alkaline conditions. Using a basic mobile phase (pH 9.2), the three aldohexoses or the three aldopentoses were simultaneously enantioseparated, respectively, but all the six monosaccharides could not be simultaneously enantioseparated. A large amount of nonreacted L-TrpNH2 was detected after the derivatized monosaccharides. In order to widen the separation window, a large portion of nonreacted L-TrpNH2 could be eliminated by liquid-liquid extraction with ethylacetate, and elution order of the derivatized monosaccharides and nonreacted L-TrpNH2 was found to be reversed using a neutral mobile phase. All of the six monosaccharides were simultaneously enantioseparated by reversed phase high-performance liquid chromatography (HPLC) using InertSustainSwift C18 column (4.6 mm i.d. × 150 mm) and a mobile phase containing 180 mM phosphate buffer (pH 7.6), 1.5 mM butylboronic acid, and 5% acetonitrile at 40 °C. Nomenclature of D and L for monosaccharides is based on the configurations of the asymmetric C4 center for aldopentoses and C5 center for aldohexoses. It was found that the enantiomer elution order of these six monosaccharides and fucose in the proposed method conformed to be the absolute configuration of the C2 center.


Subject(s)
Boron Compounds/chemistry , Chromatography, High Pressure Liquid/methods , Chromatography, Reverse-Phase/methods , Monosaccharides/isolation & purification , Tryptophan/analogs & derivatives , Arabinose/isolation & purification , Galactose/isolation & purification , Glucose/isolation & purification , Hydrogen-Ion Concentration , Liquid-Liquid Extraction , Monosaccharides/chemistry , Ribose/isolation & purification , Stereoisomerism , Tryptophan/chemistry , Xylose/isolation & purification
16.
Chirality ; 27(5): 353-7, 2015 May.
Article in English | MEDLINE | ID: mdl-25847749

ABSTRACT

Readily available L-tartaric acid, which is a bidentate ligand with two chiral centers forming a seven-membered chelate ring, was applied to the chiral ligand for the chiral nuclear magnetic resonance (NMR) shift reagent of samarium(III) formed in situ. This simple method does not cause serious signal broadening in the high magnetic field. Enantiomeric (13)C and (1)H NMR signals and enantiotopic (1)H NMR signals of α-amino acids were successfully resolved at pH 8.0 and the 1:3 molar ratio of Sm(NO3)3:L-tartaric acid. It is elucidated that the enantiomeric signal resolution is attributed to the anisotropic magnetic environment for the enantiomers induced by the chiral L-tartarato samarium(III) complex rather than differences in stability of the diastereomeric substrate adducts. The present (13)C NMR signal resolution was also effective for the practical simultaneous analysis of plural kinds of DL-amino acids.


Subject(s)
Amino Acids/chemistry , Magnetic Resonance Spectroscopy/methods , Nitrates/chemistry , Samarium/chemistry , Tartrates/chemistry , Stereoisomerism
17.
Anal Sci ; 30(7): 739-43, 2014.
Article in English | MEDLINE | ID: mdl-25007933

ABSTRACT

Three reducing monosaccharides (glucose; Glc, galactose; Gal, and mannose; Man) were derivatized with L-tryptophan (L-Trp) under alkaline conditions. The DL-Gal and DL-Man derivatives were chirally resolved by HPLC with an acidic mobile phase, but the DL-Glc derivative was not. All of the three DL-monosaccharide derivatives were simultaneously enantioseparated using HPLC with a SunShell RP-AQUA column (C28) and a basic mobile phase. The optimum mobile phase conditions consisted of 5 mM phosphate and 25 mM tetraborate buffer (pH 9.6) at 20°C. With this system, resolution of D- and L-isomers of the Glc, Gal and Man derivatives were approximately 1.7, 2.2 and 2.4, respectively. When the three monosaccarides were derivatized with L-phenylalanine instead of L-Trp, DL-Gal and DL-Man were enantioseparated under both acidic and basic conditions, but DL-Glc was not. It was observed that enantiomer elution orders of the three monosaccharides derivatized with L-Trp were reasonably reversed when derivatized with D-Trp. It was also revealed that borate anions were required for simultaneous enantioseparation with HPLC.


Subject(s)
Monosaccharides/chemistry , Tryptophan/chemistry , Boranes/chemistry , Chromatography, High Pressure Liquid , Galactose/chemistry , Hydrogen-Ion Concentration , Mannose/chemistry , Monosaccharides/isolation & purification , Spectrophotometry, Ultraviolet , Stereoisomerism , Tryptophan/isolation & purification
18.
Electrophoresis ; 34(9-10): 1327-33, 2013 May.
Article in English | MEDLINE | ID: mdl-23423790

ABSTRACT

The content of α-hydroxy acids and their enantiomers can be used to distinguish authentic and adulterated fruit juices. Here, we investigated the use of ligand exchange CE with two kinds of central metal ion in a BGE for the simultaneous determination of enantiomers of dl-malic, dl-tartaric and dl-isocitric acids, and citric acid. Ligand exchange CE with 100 mM d-quinic acid as a chiral selector ligand and 10 mM Cu(II) ion as a central metal ion could enantioseparate dl-tartaric acid but not dl-malic acid or dl-isocitric acid. Addition of 1.8 mM Sc(III) ion to the BGE with 10 mM Cu(II) ion to create a dual central metal ion system permitted the simultaneous determination of these α-hydroxy acid enantiomers and citric acid. The proposed ligand exchange CE was thus well suited for detecting adulteration of fruit juices.


Subject(s)
Beverages/analysis , Food Contamination/analysis , Fruit/chemistry , Hydroxy Acids/analysis , Citric Acid/analysis , Electrophoresis, Capillary , Isocitrates/analysis , Ligands , Malates/analysis , Metals/chemistry , Stereoisomerism , Tartrates/analysis
19.
Electrophoresis ; 33(18): 2920-4, 2012 Sep.
Article in English | MEDLINE | ID: mdl-22930546

ABSTRACT

Using two kinds of central metal ions in a background electrolyte, ligand exchange CE was investigated for the simultaneous enantioseparation of dl-malic, dl-tartaric, and dl-isocitric acids. Ligand exchange CE with 100 mM d-quinic acid as a chiral selector ligand and 10 mM Cu(II) ion as a central metal ion could enantioseparate dl-tartaric acid but not dl-malic acid or dl-isocitric acid. A dual central metal ion system containing 0.5 mM Al(III) ion in addition to 10 mM Cu(II) ion in the background electrolyte enabled the simultaneous enantioseparation of the three α-hydroxy acids. These results suggest that the use of a dual central metal ion system can be useful for enantioseparation by ligand exchange CE.


Subject(s)
Electrophoresis, Capillary/methods , Hydroxy Acids/isolation & purification , Metals, Heavy/chemistry , Acids, Acyclic/chemistry , Acids, Acyclic/isolation & purification , Electrophoresis, Capillary/instrumentation , Hydroxy Acids/chemistry , Ligands , Stereoisomerism
20.
Electrophoresis ; 33(3): 523-7, 2012 Feb.
Article in English | MEDLINE | ID: mdl-22287180

ABSTRACT

The mechanism of change in the enantiomer migration order (EMO) of tartarate on ligand exchange CE with Cu(II)- and Ni(II)-D-quinic acid systems was investigated thoroughly by circular dichroism (CD) spectropolarimetry. The (13) C NMR spectra of solutions containing D-quinate (pH 5.0) with Cu(II) or Ni(II) revealed the coordination of carboxylate and hydroxyl groups on D-quinate. The D-quinic acid concentration dependence of the CD spectra at a fixed Cu(II) concentration at pH 5.0 indicates that the 1:1, 1:2 and 1:3 Cu(II)-D-quinate complexes were formed with an increase in the concentration of D-quinic acid. The CD spectral behavior revealed that D-tartarate is selectively coordinated to the 1:1 complex to give the 1:1:1 Cu(II)-D-quinate-D-tartarate ternary complex while L-tartarate is selectively bound to the 1:2 and 1:3 complexes to form the 1:2:1 ternary complex. In the Ni(II)-D-quinic acid system, it became apparent that the 1:2 Ni(II)-D-quinate complex is mainly formed in the wide range of D-quinic acid concentration at pH 5.0 and D-tartarate is selectively coordinated to the 1:2 complex to form the 1:2:1 ternary complex. The change in EMO of tartarate on ligand exchange CE was explainable by the change in coordination selectivity for D- and L-tartarates in the Cu(II)- and Ni(II)-D-quinic acid systems depending on the compositions of the complexes formed in BGE.


Subject(s)
Copper/chemistry , Electrophoresis, Capillary/methods , Nickel/chemistry , Quinic Acid/chemistry , Tartrates/chemistry , Cations, Divalent/chemistry , Circular Dichroism , Ligands , Nuclear Magnetic Resonance, Biomolecular , Stereoisomerism , Tartrates/analysis
SELECTION OF CITATIONS
SEARCH DETAIL
...